You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/324511682

Spherical α-MnO2 Supported on N-KB as Efficient Electrocatalyst for Oxygen


Reduction in Al–Air Battery

Article in Materials · April 2018


DOI: 10.3390/ma11040601

CITATIONS READS

59 2,320

6 authors, including:

Guangli Li Quanguo He
Hunan University of Technology Hunan University of Technology
110 PUBLICATIONS 4,526 CITATIONS 178 PUBLICATIONS 7,400 CITATIONS

SEE PROFILE SEE PROFILE

Jun Liu Fuzhi Li


Hunan University of Technology Hunan University of Technology
105 PUBLICATIONS 5,169 CITATIONS 12 PUBLICATIONS 575 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Jun Liu on 29 April 2018.

The user has requested enhancement of the downloaded file.


materials
Article
Spherical α-MnO2 Supported on N-KB as Efficient
Electrocatalyst for Oxygen Reduction in
Al–Air Battery
Kui Chen, Mei Wang, Guangli Li, Quanguo He, Jun Liu * and Fuzhi Li * ID

Hunan Key Laboratory of Biomedical Nanomaterials and Devices, College of Life Sciences and Chemistry,
Hunan University of Technology, Zhuzhou 412007, China; chenkui201802@163.com (K.C.);
wangm20140900@126.com (M.W.); guangli010@163.com (G.L.); hequanguo@126.com (Q.H.)
* Correspondence: junliu@hut.edu.cn (J.L.); lifuzhi@hut.edu.cn (F.L.)

Received: 22 March 2018; Accepted: 11 April 2018; Published: 13 April 2018 

Abstract: Traditional noble metal platinum (Pt) is regarded as a bifunctional oxygen catalyst due to
its highly catalytic efficiency, but its commercial availability and application is often restricted by
high cost. Herein, a cheap and effective catalyst mixed with α-MnO2 and nitrogen-doped Ketjenblack
(N-KB) (denoted as MnO2 -SM150-0.5) is examined as a potential electrocatalyst in oxygen reduction
reactions (ORR) and oxygen evolution reactions (OER). This α-MnO2 is prepared by redox reaction
between K2 S2 O8 and MnSO4 in acid conditions with a facile hydrothermal process (named the SM
method). As a result, MnO2 -SM150-0.5 exhibits a good catalytic performance for ORR in alkaline
solution, and this result is comparable to a Pt/C catalyst. Moreover, this catalyst also shows superior
durability and methanol tolerance compared with a Pt/C catalyst. It also displays a discharge
voltage (~1.28 V) at a discharge density of 50 mA cm−2 in homemade Al–air batteries that is higher
than commercial 20% Pt/C (~1.19 V). The superior electrocatalytic performance of MnO2 -SM150-0.5
could be attributed to its higher Mn3+ /Mn4+ ratio and the synergistic effect between MnO2 and
the nitrogen-doped KB. This study provides a novel strategy for the preparation of an MnO2 -based
composite electrocatalyst.

Keywords: manganese dioxide; electrocatalyst; oxygen reduction reaction; Al-air battery; impedance;
methanol tolerance

1. Introduction
The efficiency of oxygen reduction reactions (ORR) and oxygen evolution reactions (OER) are
critical to the energy conversion efficiency of metal air batteries because of their sluggish reaction
kinetics [1]. In recent years, Pt-based materials have been developed as typical catalysts for ORR/OER
with high catalytic activity [2–4]. However, their widespread application in commerce is seriously
hindered, because the noble metal Pt is very expensive, and its reserves in earth are scarce. Therefore,
in recent years, non-precious metal-based materials were extensively studied for developing a
comparable candidate for Pt-based catalysts. Especially, a manganese-based material (such as
manganese oxide) with satisfying catalytic activity has been developed recently, because it possesses a
lot of advantages, including cheapness, abundance, environmental friendliness, structural flexibility,
and bifunctional catalytic activity for ORR/OER [5–9]. Nevertheless, many factors have been found
to play important roles in improving the catalytic activity of manganese dioxide for ORR. Firstly,
the crystalline phase of manganese dioxide is critical. It has been reported that α-MnO2 exhibits
better catalytic activity than other crystalline phases, because of its abundant di-µ-oxo bridges [10–12].
Secondly, micromorphology also has a great influence on its performance. Previous studies had shown

Materials 2018, 11, 601; doi:10.3390/ma11040601 www.mdpi.com/journal/materials


Materials 2018, 11, 601 2 of 14

that metal oxides with nanostructures exhibited good electrocatalytic performance because of their
relatively large surface area and big pore volume, which exposes more active sites and facilitates
full contact with electrolyte [13–16]. Thirdly, Mn3+ is believed to favor ORR/OER due to the single
electron occupation in σ*-orbital (eg ). Therefore, more content of Mn3+ in MnO2 could promote its
electrocatalytic performance [17,18].
Although manganese-based materials have good electrocatalytic performance as reported,
the superiority of synergic catalysts cannot be neglected. Recently, the research emphasis of manganese
oxide has focused on ion doping and its composition with other materials, and results indicate that it
has better catalytic property than bare manganese oxide. For example, Fe (or Co) ion-doped MnO2
nanosheets (MONSs) grown on the internal surface of macroporous carbon showed improved ORR
catalytic activity compared with the un-doped one, because of the co-electrocatalytic function of MnO2
and Fe (or Co ion) [19]. Moreover, the catalyst of Mn2 O3 -doped MnO supported by reduced grapheme
oxide (rGO) was proved to have a better ORR catalytic performance and stronger stability than that
of pure MnO. It was believed that the coexisted metal oxide with different valences and rGO had
promoted the catalytic performance [20]. In addition, carbon is one of the most important materials
for electron transfer, and it could improve the catalytic activity. The coupling of MnO2 with carbon
materials may improve its catalytic activities.
Herein, MnO2 spheres are synthesized by the redox reactions between K2 S2 O8 and MnSO4 in acid
conditions (denoted as the SM method). These MnO2 spheres mixed with nitrogen-doped Ketjenblack
(N-KB) are used as a catalyst for ORR/OER application. This study examines the morphology,
structure, and electrochemical properties of MnO2 samples by scanning electron microscope (SEM),
X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and electrochemical testing.
The relationships between the synthesized conditions and electrocatalytic activity of these MnO2 -N-KB
catalysts are also discussed. Finally, the proposed enhanced catalytic mechanism of these composited
catalysts is investigated by controlling the Mn3+ content in MnO2 . This work offers a new strategy for
the scalable preparation of more efficient MnO2 bifunctional oxygen catalysts for ORR and OER.

2. Experiment

2.1. Preparation of MnO2


First, 1.69 g of manganese sulfate monohydrate, 1.74 g of potassium sulfate, and 2.71 g of
potassium persulfate were dissolved into 80 mL of deionized water, agitating at least 5 min to form
a homogenous aqueous solution. Then, 4 mL of 37% hydrochloric acid was added into the aqueous
solution. After agitating another 5 min, the resulting solution was transferred into a 100-mL Teflon
lined stainless steel shell autoclave, heated to 120 ◦ C in an oven, and kept at the temperature for 12 h.
After it was cooled to room temperature naturally, the as-obtained product was filtrated under vacuum
with a membrane of 0.15-µm pore diameter and dried overnight at 80 ◦ C. The obtained product was
denoted as MnO2 -SM120-12 (S: potassium persulfate; M: manganese sulfate).
For comparison, the control samples, such as MnO2 -SM150-0.5 and MnO2 -SM120-0.5, were also
prepared. For the samples of MnO2 -SM150-0.5, the synthesis process was the same with that of
the MnO2 -SM120-12, but the reaction temperature and reaction time were 150 ◦ C and 30 min,
respectively. For the samples of MnO2 -SM120-0.5, the synthesis process was the same with that of the
MnO2 -SM120-12, but the reaction temperature and reaction time were 120 ◦ C and 30 min, respectively.

2.2. Preparation of N-KB


First, 0.2 g of ketjenblack (KB) and 1.2 g of melamin were dispersed in 80 mL of deionized water
by ultrasonic treatment for 30 min. Then, the resulting solution was enclosed into a 100-mL Teflon
lined stainless steel shell autoclave, heated to 120 ◦ C in an oven, and kept at the temperature for 24 h.
After it was cooled down to room temperature naturally, the as-obtained product was filtrated under
vacuum with a membrane of 0.15-µm pore diameter, and dried overnight at 80 ◦ C. After being ground
Materials 2018, 11, 601 3 of 14

by agate mortar for more than 10 min, the resulting powder was transferred to a piece of porcelain
boat, which was then covered with another piece of porcelain boat, and further wrapped by copper
foil. The treated porcelain boat was placed into a tube furnace, and then heated to 650 ◦ C for 2 h at a
heating rate of 5 ◦ C min−1 in argon flow. After that, it was naturally cooled down to room temperature,
and the as-prepared sample was denoted as N-KB.

2.3. Characterization
The morphologies of the as-prepared catalysts were characterized by using the scanning electron
microscope (FIB 600i, FEI, Hillsboro, OR, USA). The structures of these samples were carried out
by X-ray diffraction (XRD, Rigaku D/Max 2550, Tokyo, Japan) with Cu-Kα radiation (λ = 1.5406 Å).
The elemental and valence state analysis were characterized by X-ray photoelectron spectroscopy
(XPS, K-Alpha1063 spectrometer, Thermo Scientific Co., Waltham, MA, USA).

2.4. Electrochemical Measurements


For electrochemical measurements, 2 mg of as-prepared MnO2 and 4 mg of N-KB were dispersed
in 950 µL of anhydrous ethanol by sonication for 20 min. Then, 50 µL of Nafion solution (5 wt %)
was added and sonicated for another 20 min to get a homogeneous catalytic ink. Then, 8 µL
of the ink was loaded onto the surface of glassy carbon disk (the diameter is 5 mm, homemade
electrode), and the catalyst loading amount was 0.0815 mg cm−2 (calculated by the mass of MnO2 ).
For comparison, the commercial 20 wt % Pt/C (Johnson Matthey, Royston, UK) was also prepared
with the same method.
Linear sweep voltammetry (LSV) and cyclic voltammetry (CV) ORR were performed using a
RDE (rotating disk electrode) as working electrode, a double fluid boundary Ag/AgCl electrode as the
reference electrode, and a platinum wire as the counter electrode in 0.1-M of KOH solution saturated
with oxygen on a CHI760E electrochemical workstation. All of the potentials were finally converted
to the values versus reversible hydrogen electrode. The ORR catalytic stabilities were evaluated by
half-wave potential decay (∆E1/2 ) before and after the accelerated durability test (ADT). The ADT was
performed by using these catalysts in ORR for 5000 cycles. These experiments were carried out in
O2 -saturated 0.1 M of KOH solution at room temperature, and the voltage was selected from 0.57 V
to 0.82 V (versus (Reversible Hydrogen Electrode) RHE) with a scan rate of 100 mV s−1 . Methanol
tolerance testing of catalysts was carried out in O2 -saturated 0.1 M of KOH and 1.0 M of CH3 OH-mixed
electrolyte [21].
To further verify the ORR mechanism, the RRDE (rotating ring disk electrode) technique was
used, and the peroxide percentage and the electron transfer number were calculated based on the
equations, which were given as follows [22]:

Ir/N
HO2− (%) = 200 × (1)
Id + Ir/N

Id
n = 4× (2)
Id + Ir/N
where Id represents the disk current, Ir represents the ring current, N represents the current collection
efficiency of the Pt ring (0.37), and n means the electron transfer number [22,23].
The OER activities of the as-prepared samples were also carried out by RDE experiments at a
scan rate of 10 mV s−1 with a rotation speed of 1600 rpm. The current density was operated from 0 to
16.5 mA cm−2 . The long-term durability OER measurements of the catalysts were performed by using
chronopotentiometry. The tests were conducted at a current density of 10 mA cm−2 , and the test time
was 8000 s. Finally, the EIS testes were scanned in the frequency range of 105 –0.1 Hz at 1.665 V (versus
RHE) with the amplitude of 5 mV in 0.1 M of KOH solution [24].
Materials 2018, 11, 601 4 of 14

2.5. Electrochemical Test of Al–Air Batteries


For an Al–air full batteries test, a polished aluminum stripe was used as anode, and 6 mol L−1 of a
KOH solution containing 0.01 mol L−1 of Na2 SnO3 , 0.0075 mol L−1 of ZnO, and 0.0005 mol L−1 of In(OH)3
was used as electrolyte. The air electrodes were composed of a gas diffusion layer, a foam nickel
current collector, and a catalytic layer. The catalytic layer was fabricated as follows. The as-prepared
catalysts (10 mg), N-KB (30 mg), and the 60 wt % polytetrafluoroethylene (PTFE) aqueous solution
(~50 mg) were mixed and agitated continuously until a paste appeared. Then, this paste was rolled
with a glass rod until it turned into a 2 cm × 2 cm film. In the end, the film and the gas diffusion
layer were pressed onto the two sides of nickel foam under the pressure of 10 MPa, and dried at
60 ◦ C overnight. For comparison, the air electrode using commercial 20 wt % Pt/C catalyst was also
fabricated with the same method. The full battery performance was measured with a Neware Battery
Testing System (Shenzhen, China). A homemade electrochemical cell was used for Al–air battery
measurements, with the net volume size of 50 mm × 32 mm × 50 mm, and an air hole with a diameter
of 10 mm was used for the test [22,25].

3. Results and Discussion

3.1. Micromorphological and Microstructural Properties


The micromorphology of the synthesized samples was characterized by SEM. Figure 1 shows
the SEM images of MnO2 -SM120-12 (a–c), MnO2 -SM120-0.5 (d–f), and MnO2 -SM150-0.5 (g–i). As we
can see from these images, all three samples display spherical morphology. The average size of the
MnO2 -SM120-12 sample was ~5.0 µm, and these spheres were composed of nanorods with average
diameters of 50 nm. The average diameter of the MnO2 -SM150-0.5 spherical particles was ~4.2 µm,
and are composed of nanorods with an average diameter of 27 nm. The spherical MnO2 -SM120-0.5
sample shows an average diameter of ~2.9 µm, and is composed with nanorods (whose average
diameter is about 20 nm). The MnO2 -SM150-0.5 shows a smaller size than MnO2 -SM120-12, which is
mainly because of the higher hydrothermal reaction temperature and shorter hydrothermal reaction
time. Although the manganese dioxide formed at a higher reaction speed during the production
process, the samples with a smaller particle size could be produced in less reaction time [10]. However,
the MnO2 -SM150-0.5 sample shows a larger size than MnO2 -SM120-0.5; this should be attributed to
the effect of the higher reacting temperature, which causes a quicker reaction speed, and the quicker
speed leads to a bigger size in the same reaction time.
Materials 2018, 11, 601 5 of 14

Figure 1. The SEM images of MnO2 -SM120-12 (a–c); MnO2 -SM120-0.5 (d–f); and MnO2 -SM150-0.5 (g–i).

The structural characterization of MnO2 -SM120-12 (green line), MnO2 -SM120-0.5 (blue line),
and MnO2 -SM150-0.5 (orange line), were carried out and compared by XRD. As shown in these XRD
patterns (Figure 2), the diffraction peaks located at the 12.8, 18.1, 25.8, 28.8, 37.5, 42.0, 49.9, 56.9, 60.3,
65.1, 69.7 and 72.7 positions in all three samples belonged to the (110), (200), (220), (310), (211), (301),
(411), (431), (521), (002), (541) and (312) facets, respectively. These sharp peaks indicated the good
crystallization property of the samples, and there no other diffraction peaks were observed in these
patterns, further indicating the good crystallization property of these samples. The standard card
of α-MnO2 (PDF#44-0141) was presented for comparison, and these diffraction peaks matched well
with this standard card. It revealed that these as-synthesized MnO2 particles are α-MnO2 particles.
Obviously, the diffraction peaks intensity of MnO2 -SM150-0.5 (orange line) were inferior to those of the
other two samples; this is probably because more defects existed in this sample, which could improve
its catalytic activities. Moreover, it has been generally accepted that α-MnO2 exhibits better catalytic
activity than other crystalline phases because of its abundant di-µ-oxo bridges [11]. Thus, the α-MnO2
particles prepared here could be used for high-performance catalytic application.
Materials 2018, 11, 601 6 of 14

Figure 2. XRD patterns of MnO2 -SM120-12 (green line), MnO2 -SM120-0.5 (blue line), and MnO2 -SM150-0.5
(orange line), the standard PDF card of MnO2 (PDF#44-0141) is carried out for comparison.

XPS spectra were carried out for further elemental and valence state analysis of MnO2 -SM120-12,
MnO2 -SM120-0.5, and MnO2 -SM150-0.5 samples. As shown in Figure 3, the high-resolution XPS
spectra of Mn 2p for MnO2 -SM120-12 (Figure 3a), MnO2 -SM120-0.5 (Figure 3b), and MnO2 -SM150-0.5
(Figure 3c) are presented, and four peaks located at 642.30, 643.25, 653.80, and 654.80 eV were obtained
by peak-differentiating technique. These peaks were assigned to the Mn3+ (2p3/2 ), Mn4+ (2p3/2 ),
Mn3+ (2p1/2 ), and Mn4+ (2p1/2 ) species, respectively [6,26,27]. Moreover, based on the XPS results,
the perk areas of Mn3+ and Mn4+ in each sample are presented in Table 1, and the ratio values of
Mn3+ /Mn4+ (Figure 3d) for MnO2 -SM120-12, MnO2 -SM120-0.5, and MnO2 -SM150-0.5 were calculated
as 0.813, 0.512 and 0.965, respectively. As shown in Figure 3d, obviously, the Mn3+ /Mn4+ ratio in
MnO2 -SM150-0.5 (0.965) was higher than that in MnO2 -SM120-12 (0.813), which should be ascribed
to the higher reacting temperature that causes the faster reaction rate. It is believed that the faster
reaction rate could cause more Mn3+ content in MnO2 . A higher content of Mn3+ in MnO2 can
lead to a better electrocatalytic performance, due to the single electron occupation in the σ*-orbital
(eg ) of Mn3+ [17,18,28]. In addition, a shorter reacting time could produce MnO2 -SM150-0.5 with a
smaller size (Figure 1), which would impact its electrocatalytic performance. The Mn3+ /Mn4+ ratio
in MnO2 -SM150-0.5 (0.965) was also higher than that in MnO2 -SM120-0.5 (0.512), which should be
because the faster reaction rate could cause more Mn3+ content in MnO2 [10].
Materials 2018, 11, 601 7 of 14

Figure 3. XPS spectra of Mn 2p for MnO2 -SM120-12 (a); MnO2 -SM120-0.5 (b); MnO2 -SM150-0.5 (c);
and the Mn3+ /Mn4+ values of the three samples (d).

Table 1. The XPS dates of Mn 2p for MnO2 -SM120-12; MnO2 -SM120-0.5; MnO2 -SM150-0.5; and the
corresponding perk areas and Mn3+ /Mn4+ values.

Samples Species Peak Position (eV) Peak Area Mn3+ /Mn4+


2p 3/2 Mn3+ 642.60 36429.72
2p 3/2 Mn4+ 643.60 33877.38
MnO2 -SM150-0.5 0.965
2p 1/2 Mn3+ 654.00 18763.28
2p 1/2 Mn4+ 654.90 23322.43
2p 3/2 Mn3+ 642.30 20298.63
2p 3/2 Mn4+ 643.25 44859.42
MnO2 -SM120-0.5 0.512
2p 1/2 Mn3+ 653.80 14298.36
2p 1/2 Mn4+ 654.80 22753.79
2p 3/2 Mn3+ 642.30 29778.53
2p 3/2 Mn4+ 643.25 45449.07
MnO2 -SM120-12 0.813
2p 1/2 Mn3+ 653.80 23045.48
2p 1/2 Mn4+ 654.80 19499.66

3.2. ORR Activity and Stability


The LSV curves for the ORR of MnO2 -SM120-12, MnO2 -SM120-0.5, and MnO2 -SM150-0.5
are shown in Figure 4a; these measurements were carried out in 0.1 M of KOH solution at
a rotating speed of 1600 rpm. The MnO2 -SM150-0.5 sample showed a better ORR catalytic
performance than MnO2 -SM120-12 at the same test conditions, because the size of the particles
Materials 2018, 11, 601 8 of 14

in the MnO2 -SM150-0.5 sample was smaller, and it contained more Mn3+ in comparison with the
MnO2 -SM120-12 sample [28,29]. It is clear that MnO2 -SM150-0.5 also exhibited a better ORR catalytic
performance than the MnO2 -SM120-0.5 samples, mainly because of the higher Mn3+ content. As shown
in Figure 4b, MnO2 -SM150-0.5 (supported on N-KB) exhibited a much better ORR catalytic performance
than bare N-KB, with a half-wave potential of 0.76 V and a limiting current density of 6.0 mA cm−2 .
This phenomenon should be attributed to the synergetic catalytic activity of α-MnO2 and N-KB.
It is believed that the intrinsically abundant di-µ-oxo bridges in α-MnO2 could facilitate the ORR
process [7,12,19,28]. As observed, the limiting current density (6.0 mA cm−2 ) of MnO2 -SM150-0.5
(supported on N-KB) was higher than Pt/C (~5.0 mA cm−2 ), despite its lower half-wave potential
(0.76 V) compared with Pt/C (~8.2 V). As shown in Figure 4c, the average n value of MnO2 -SM150-0.5
was 3.85 (from 3.78 to 3.95), confirming a four-electron (4e− ) oxygen reduction mechanism.

Figure 4. (a) Oxygen reduction reaction (ORR) linear sweep voltammetries (LSVs) of MnO2 -SM120-12,
MnO2 -SM120-0.5, and MnO2 -SM150-0.5 in 0.1 M of KOH solution at a scan rate of 10 mV s−1 with a
rotation speed of 1600 rpm; (b) ORR LSVs of MnO2 -SM120-0.5, N-KB, and 20% JM. Platinum (Pt)/C in
0.1 M of KOH solution at a scan rate of 10 mV s−1 with a rotation speed of 1600 rpm; (c) Percentage of
peroxide (HO2− %) and the electron transfer number (n) of MnO2 -SM150-0.5.

The catalytic stabilities of the MnO2 -SM150-0.5 and Pt/C samples in ORR were evaluated by the
half-wave potential decay (∆E1/2 ) before and after the accelerated durability test (ADT) [30]. The ADT
was performed by using these catalyst in an ORR for 5000 cycles. These experiments were carried out
in an O2 -saturated 0.1 M of KOH solution at room temperature, and the voltage was selected from
0.57 V to 0.82 V (versus RHE) with a scan rate of 100 mV s−1 . As shown in Figure 5a, the half-wave
potential of the MnO2 -SM150-0.5 sample (supported on N-KB) exhibited a negative shift of ~33 mV
after 5000 cycles; this result is slightly higher than that of Pt/C (~22 mV, Figure 5b), which is probably
because of the inferior electroconductibility of MnO2 compared with the noble metal Pt.
Materials 2018, 11, 601 9 of 14

Figure 5. LSV curves of MnO2 -SM150-0.5 (a) and Pt/C (b) before and after the accelerated durability
test (ADT). The ADT was performed by subjecting the catalyst to 5000 circles from 0.57 V to 0.82 V
(vs. RHE) in an O2 -saturated 0.1 M of KOH solution at room temperature at a scan rate of 100 mV s−1 .

3.3. OER Activity and Stability


In addition, the OER activities of three as-prepared samples (MnO2 -SM120-12, MnO2 -SM120-0.5,
and MnO2 -SM150-0.5) were carried out for further comparing their electrocatalytic performances,
and they were tested by RDE experiments at the scan rate of 10 mV s−1 with a rotation speed of
1600 rpm. Generally, OER activities are judged by the potential at the current density from 0 mA cm−2
to 16.5 mA cm−2 [31]. As presented in Figure 6a, MnO2 -150-0.5 showed a more negative shift than the
MnO2 -120-12 and MnO2 -SM120-0.5 samples when the current density increased from 0 mA cm−2 to
16.5 mA cm−2 , which means that the MnO2 -15-0.5 could catalyze OER at a lower overpotential than
the MnO2 -120-12 and MnO2 -SM120-0.5 samples. In other words, the MnO2 -150-0.5 sample exhibited
much better OER kinetic behavior than the MnO2 -120-12 and MnO2 -SM120-0.5 samples. Similar to
the ORR, the higher content of Mn3+ and smaller size of the α-MnO2 nanorods also played a vital
role on the OER. Moreover, the stability tests of the OER activity of the MnO2 -SM150-0.5 and Pt/C
samples are presented. The long-term durability measurements of the catalysts were performed by
using chronopotentiometry. The tests were conducted at a current density of 10 mA cm−2 , and the test
time was 8000 s. As presented in Figure 6b, after reaction for 8000 s, MnO2 -SM150-0.5 showed a good
stability in OER, but slightly not as good as Pt/C, indicating that the practical catalytic performance of
MnO2 -SM150-0.5 needs further improvement to replace the Pt/C.

Figure 6. The oxygen evolution reaction (OER) performances of MnO2 -150-0.5 and MnO2 -120/12 are
evaluated by LSVs in 0.1 M of KOH solution at a scan rate of 10 mV s−1 with a rotation speed of
1600 rpm (a); the stability test of OER activity of MnO2 -SM150-0.5 and Pt/C samples in 1.0 M of KOH
solution at a current density of 10 mA cm−2 for 8000 s (b).
Materials 2018, 11, 601 10 of 14

3.4. EIS Performance


The charge transfer efficiency of the catalyst plays an important role in the OER process, as the high
electron transfer efficiency could indicate the high catalytic activity. The electrochemical impedance
spectroscopy (EIS) is a good method for understanding charge transfer efficiency, because the arc
radius size of the EIS curve could indicate the value of electrical resistance. The lower resistance of
the sample implies the high electroconductibility. Thus, the EIS method is used for deep insights
into the OER process. The EIS tests were scanned in the frequency range of 105 –0.1 Hz at 1.665 V
(versus RHE) with the amplitude of 5 mV in 0.1 M of KOH solution [17,32]. The Nyquist plots are
shown in Figure 7, in which the EIS data (Figure 7 left) have been fitted according to the equivalent
circuit (Figure 7 right). The equivalent circuit consisted of Rs , Rf , Rct , C, and CPE, representing the
uncompensated solution resistance, intrinsic resistance of the catalyst, charge transfer resistance,
capacitance of catalyts, and constant phase element of the double layer, respectively. All of the fitting
parameters are listed in Table 2. The Rs relating to the uncompensated solution of the MnO2 -120-0.5,
MnO2 -120-12, and MnO2 -150-0.5 samples are 55.72 Ω, 61.75 Ω, and 62.01 Ω, respectively. The Rct
relates to the reaction kinetics; MnO2 -150-0.5 exhibited a lower charge transfer resistance (146.4 Ω)
than those of the MnO2 -120-12 (257.1 Ω) and MnO2 -120-0.5 (313.5 Ω) samples. Note that it is in good
accordance with the OER performance.

Figure 7. Nyquist plots of MnO2 -120-0.5, MnO2 -120-12, and MnO2 -150-0.5 samples obtained from
electrochemical impedance spectroscopy (EIS) measurements in 0.1 M of KOH solution at 1.665 V
(vs. RHE) and the inserted is the corresponding equivalent circuit.

Table 2. Component values of fitted equivalent circuit based on the Nyquist plots.

Sample Rs (Ω) Rf (Ω) Rct (Ω) C (F) CPE-T CPE-P


MnO2 -SM150-0.5 55.72 4.77 146.4 0.001010 0.002368 0.842
MnO2 -SM120-12 61.75 8.89 257.1 0.005600 0.003926 0.692
MnO2 -SM120-0.5 62.01 9.69 313.5 0.000031 0.002031 0.895

3.5. Methanol Tolerance Performance


The methanol tolerance of the catalysts is usually used to evaluate the performance of ORR
catalysts in DMFCs (direct methanol fuel cells). The LSV and CV experimental groups (by using the
MnO2 -SM150-0.5 sample (supported on N-KB) and Pt/C samples as catalysts) and control group were
carried out in an O2 -saturated 0.1 M of KOH electrolyte for testing the methanol tolerance. These results
are presented in Figure 8. As we can see in Figure 8a, when the ORR process was carried out in an
O2 -saturated 0.1 M of KOH electrolyte with 1.0 M of methanol, the MnO2 -SM150-0.5 catalyst exhibited
excellent methanol tolerance properties, because there was no negative shift of onset potential and
no oxidation currents of methanol, but rather only a slight decrease of the limiting current density
Materials 2018, 11, 601 11 of 14

(~0.25 mA cm−2 ) (Figure 8b). However, a strong oxidation current of methanol is shown in Figure 8d
with the Pt/C sample by comparison with the background line. Moreover, a larger negative shift of
onset potential for ORR (from ~1.0 V to ~0.52 V) is observed in Figure 8c, indicating the poor methanol
tolerance of the Pt/C sample in comparison with the MnO2 -SM150-0.5 catalyst [9].

Figure 8. LSV curves of MnO2 -SM150-0.5 (a) and Pt/C (c) in an O2 -saturated 0.1 M of KOH electrolyte
with (purple line) and without (blue line) 1.0 M methanol; cyclic voltammetry (CV) curves of
MnO2 -SM150-0.5 (b) and Pt/C (d) at 20 mV s−1 in an O2 -saturated 0.1 M of KOH electrolyte with
(red line) and without (black line) 1.0 M methanol.

3.6. Application in Al-Air Battery


For further evaluating the practical catalytic performance of the MnO2 -SM150-0.5 sample in
an Al–air battery, cell voltages at various current densities and constant current discharge tests
were carried out. The commercial Pt/C sample was also investigated for comparison. As the
results show in Figure 9a, overall, the cell polarization curve with the MnO2 -SM150-0.5 sample
was better than that of Pt/C. Specifically, when the discharge current density was lower than
150 mA cm−2 , the cell voltages of MnO2 -SM150-0.5 were higher than those of the Pt/C sample.
However, the cell voltages of the MnO2 -SM150-0.5 sample were almost equal to those of Pt/C at the
range of 150 mA cm−2 –180 mA cm−2 . As shown in Figure 9b, MnO2 -SM150-0.5 showed a discharge
voltage platform of ~1.24 V, which was slightly higher than that of Pt/C (~1.19 V) at the end of the
discharge test with a constant current density of 50 mA cm−2 in homemade Al–air batteries. However,
it can be observed that the MnO2 -SM150-0.5 sample took about 2 h to achieve the smoothing discharge
voltage platform, which was longer than that of Pt/C (less than 1 h), indicating that the practical
catalytic performance of MnO2 -SM150-0.5 needs further improvement to replace the Pt/C.
Materials 2018, 11, 601 12 of 14

Figure 9. (a) Polarization curves of Al–air batteries and (b) discharge curves at a constant current
density of 50 mA cm−2 .

4. Conclusions
In this work, three kinds of α-MnO2 microspheres composed with nanorods were synthesized
in acid conditions using K2 S2 O8 and MnSO4 as raw materials by a facile hydrothermal process.
The influences of Mn3+ content on the electrocatalytic activity of ORR/OER were also studied.
These results demonstrated that catalysts with more Mn3+ content play an important role in
electrocatalytic application. Especially, the MnO2 -SM150-0.5 sample with higher Mn3+ content
showed a better electrocatalytic performance than the MnO2 -SM120-0.5 and MnO2 -SM120-12 samples.
The half-wave potential (E1/2 ) of the MnO2 -SM150-0.5/N-KB sample was 0.76 V (versus RHE), and the
limiting current density was about 6.0 mA cm−2 . This result could be comparable to those of Pt/C
(0.82 V and ~5.0 mA cm−2 , respectively). Moreover, the MnO2 -SM150-0.5 sample showed an excellent
methanol tolerance compared to the Pt/C sample. In addition, the MnO2 -SM150-0.5 sample exhibited
good ORR catalytic stability; as its half-wave potential only negatively shifted ~33 mV after 5000
cycles. Besides, the MnO2 -SM150-0.5 sample exhibited a higher discharge voltage (1.28 V) at a density
of 50 mA cm−2 than the Pt/C catalyst (1.19 V) when used in homemade Al–air batteries as cathode
catalysts. Thus, this strategy for the preparation of α-MnO2 could provide a scalable preparation
method for significant ORR/OER application.

Acknowledgments: This work was financially supported by the National Nature Science Foundation of China
(Grant No. 61703152), Hunan Provincial Natural Science Foundation (2018JJ3134) and Zhuzhou Science and
Technology Plans (201707201806), the Undergraduates Innovation Funds of Hunan University of Technology
(HUT2018), and the Undergraduates Innovation Funds of College of Life Science and Chemistry (CLSC2018).
Author Contributions: Jun Liu and Fuzhi Li conceived and designed the experiments; Kui Chen and Mei Wang
performed the experiments; Jun Liu, Quanguo He, Guangli Li and Fuzhi Li analyzed the data; Kui Chen, Mei Wang,
Guangli Li, Quanguo He, Jun Liu and Fuzhi Li contributed reagents/materials/analysis tools; Kui Chen and
Mei Wang wrote the first draft, Jun Liu and Fuzhi Li revised and finished the paper.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Stacy, J.; Regmi, Y.N.; Leonard, B.; Fan, M. The recent progress and future of oxygen reduction reaction
catalysis: A review. Renew. Sustain. Energy Rev. 2017, 69, 401–414. [CrossRef]
2. Marković, N.M.; Schmidt, T.J.; Stamenković, V.; Ross, P.N. Oxygen reduction reaction on Pt and Pt bimetallic
surfaces: A selective review. Fuel Cells 2001, 1, 105–116. [CrossRef]
3. Li, J.; Xi, Z.; Pan, Y.-T.; Spendelow, J.S.; Duchesne, P.N.; Su, D.; Li, Q.; Yu, C.; Yin, Z.; Shen, B.; et al.
Fe stabilization by intermetallic L10-FePt and Pt catalysis enhancement in L10-FePt/Pt nanoparticles for
efficient oxygen reduction reaction in fuel cells. J. Am. Chem. Soc. 2018, 140, 2926–2932. [CrossRef] [PubMed]
Materials 2018, 11, 601 13 of 14

4. Sui, S.; Wang, X.; Zhou, X.; Su, Y.; Riffat, S.; Liu, C.-J. A comprehensive review of Pt electrocatalysts for
the oxygen reduction reaction: Nanostructure, activity, mechanism and carbon support in PEM fuel cells.
J. Mater. Chem. A 2017, 5, 1808–1825. [CrossRef]
5. Yang, S.; Wang, Z.; Cao, Z.; Mao, X.; Shi, M.; Li, Y.; Zhang, R.; Yin, Y. Facile synthesis of well dispersed spinel
cobalt manganese oxides microsphere as efficient bi-functional electrocatalysts for oxygen reduction reaction
and oxygen evolution reaction. J. Alloys Compd. 2017, 721, 482–491. [CrossRef]
6. Xue, Y.; Miao, H.; Sun, S.; Wang, Q.; Li, S.; Liu, Z. (La1−x Srx )0.98 MnO3 perovskite with A-site deficiencies
toward oxygen reduction reaction in aluminum-air batteries. J. Power Sources 2017, 342, 192–201. [CrossRef]
7. Wang, Y.; Liu, Q.; Hu, T.; Zhang, L.; Deng, Y. Carbon supported MnO2 -CoFe2 O4 with enhanced
electrocatalytic activity for oxygen reduction and oxygen evolution. Appl. Surf. Sci. 2017, 403, 51–56.
[CrossRef]
8. Wang, B.; Pan, S.; Qin, Y.; Tan, W.; Tao, Y.; Li, K.; Kong, Y. Synthesis of efficient electrocatalyst for oxygen
reduction reaction by using poly(m-phenylenediamine) as the interlayer spacer and the sources of N-doped
carbon and MnO. Synth. Met. 2017, 224, 92–98. [CrossRef]
9. Tang, Y.; Chen, T.; Guo, W.; Chen, S.; Li, Y.; Song, J.; Chang, L.; Mu, S.; Zhao, Y.; Gao, F. Reduced graphene
oxide supported MnS nanotubes hybrid as a novel non-precious metal electrocatalyst for oxygen reduction
reaction with high performance. J. Power Sources 2017, 362, 1–9. [CrossRef]
10. Zheng, X.; Yu, L.; Lan, B.; Cheng, G.; Lin, T.; He, B.; Ye, W.; Sun, M.; Ye, F. Three-dimensional radial α-MnO2
synthesized from different redox potential for bifunctional oxygen electrocatalytic activities. J. Power Sources
2017, 362, 332–341. [CrossRef]
11. Xiao, W.; Wang, D.; Lou, X.W. Shape-Controlled Synthesis of MnO2 Nanostructures with enhanced
electrocatalytic activity for oxygen reduction. J. Phys. Chem. C 2010, 114, 1430–1434. [CrossRef]
12. Meng, Y.; Song, W.; Huang, H.; Ren, Z.; Chen, S.Y.; Suib, S.L. Structure-property relationship of bifunctional
MnO2 nanostructures: Highly efficient, ultra-stable electrochemical water oxidation and oxygen reduction
reaction catalysts identified in alkaline media. J. Am. Chem. Soc. 2014, 136, 11452–11464. [CrossRef] [PubMed]
13. Ma, Y.; Wang, R.; Wang, H.; Key, J.; Ji, S. Control of MnO2 nanocrystal shape from tremella to nanobelt for
ehancement of the oxygen reduction reaction activity. J. Power Sources 2015, 280, 526–532. [CrossRef]
14. Selvakumar, K.; Thangamuthu, R.; Ganesan, K.; Murugan, P.; Rajput, P.; Jha, S.N.; Bhattacharyya, D.
Physiochemical investigation of shape-designed MnO2 nanostructures and their influence on oxygen
reduction reaction activity in alkaline solution. J. Phys. Chem. C 2015, 119, 6604–6618. [CrossRef]
15. Selvakumar, K.; Kumar, S.M.S.; Thangamuthu, R.; Kruthika, G.; Murugan, P. Development of
shape-engineered α-MnO2 materials as bi-functional catalysts for oxygen evolution reaction and oxygen
reduction reaction in alkaline medium. Int. J. Hydrogen Energy 2014, 39, 21024–21036. [CrossRef]
16. Zhang, Q.; Wang, J.; Dong, J.; Ding, F.; Li, X.; Zhang, B.; Yang, S.; Zhang, K. Facile general strategy toward
hierarchical mesoporous transition metal oxides arrays on three-dimensional macroporous foam with
superior lithium storage properties. Nano Energy 2015, 13, 77–91. [CrossRef]
17. Suntivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, J.B.; Shao-Horn, Y. A perovskite oxide optimized
for oxygen evolution catalysis from molecular orbital principles. Science 2011, 334, 1383–1385. [CrossRef]
[PubMed]
18. Suntivich, J.; Gasteiger, H.A.; Yabuuchi, N.; Nakanishi, H.; Goodenough, J.B.; Shao-Horn, Y. Design principles
for oxygen-reduction activity on perovskite oxide catalysts for fuel cells and metal–air batteries. Nat. Chem.
2011, 3, 546–550. [CrossRef] [PubMed]
19. Li, J.; Ren, Y.; Wang, S.; Ren, Z.; Yu, J. Transition metal doped MnO2 nanosheets grown on internal surface of
macroporous carbon for supercapacitors and oxygen reduction reaction electrocatalysts. Appl. Mater. Today
2016, 3, 63–72. [CrossRef]
20. Wu, X.; Gao, X.; Xu, L.; Huang, T.; Yu, J.; Wen, C.; Chen, Z.; Han, J. Mn2 O3 doping induced the improvement
of catalytic performance for oxygen reduction of MnO. Int. J. Hydrogen Energy 2016, 41, 16087–16093.
[CrossRef]
21. Wang, X.; Wang, J.; Wang, D.; Dou, S.; Ma, Z.; Wu, J.; Tao, L.; Shen, A.; Ouyang, C.; Liu, Q.; et al. One-pot
synthesis of nitrogen and sulfur co-doped graphene as efficient metal-free electrocatalysts for the oxygen
reduction reaction. Chem. Commun. 2014, 50, 4839–4842. [CrossRef] [PubMed]
Materials 2018, 11, 601 14 of 14

22. Li, F.; Fu, L.; Li, J.; Yan, J.; Tang, Y.; Pan, Y.; Wang, H. Ag/Fe3 O4 -N-doped ketjenblack carbon composite as
highly efficient oxygen reduction catalyst in al-air batteries. J. Electrochem. Soc. 2017, 164, A3595–A3601.
[CrossRef]
23. Li, J.; Chen, J.; Wang, H.; Ren, Y.; Liu, K.; Tang, Y.; Shao, M. Fe/N co-doped carbon materials with
controllable structure as highly efficient electrocatalysts for oxygen reduction reaction in Al-air batteries.
Energy Storage Mater. 2017, 8, 49–58. [CrossRef]
24. Stoerzinger, K.A.; Risch, M.; Han, B.; Yang, S.H. Recent insights into manganese oxides in catalyzing oxygen
reduction kinetics. ACS Catal. 2015, 5, 446–449. [CrossRef]
25. Liu, K.; Peng, Z.; Wang, H.; Ren, Y.; Liu, D.; Li, J.; Tang, Y.; Zhang, N. Fe3 C@Fe/N doped graphene-like
carbon sheets as a highly efficient catalyst in Al-Air batteries. J. Electrochem. Soc. 2017, 164, F475–F483.
[CrossRef]
26. Pargoletti, E.; Cappelletti, G.; Minguzzi, A.; Rondinini, S.; Leoni, M.; Marelli, M.; Vertova, A.
High-performance of bare and Ti-doped α-MnO2 nanoparticles in catalyzing the oxygen reduction reaction.
J. Power Sources 2016, 325, 116–128. [CrossRef]
27. Ozcan, S.; Tokur, M.; Cetinkaya, T.; Guler, A.; Uysal, M.; Guler, M.O.; Akbulut, H. Free standing flexible
graphene oxide + α-MnO2 composite cathodes for Li–Air batteries. Solid State Ion. 2016, 286, 34–39.
[CrossRef]
28. Wei, Z.H.; Zhao, T.S.; Zhu, X.B.; Tan, P. MnO2−x nanosheets on stainless steel felt as a carbon- and binder-free
cathode for non-aqueous lithium-oxygen batteries. J. Power Sources 2016, 306, 724–732. [CrossRef]
29. Zhang, Q.; Cheng, X.; Feng, X.; Qiu, G.; Tan, W.; Liu, F. Large-scale size-controlled synthesis of
cryptomelane-type manganese oxide OMS-2 in lateral and longitudinal directions. J. Mater. Chem. A
2011, 21, 462–465. [CrossRef]
30. Sun, S.; Zhang, G.; Geng, D.; Chen, Y.; Li, R.; Cai, M.; Sun, X. A highly durable platinum nanocatalyst for
proton exchange membrane fuel cells: Multiarmed starlike nanowire single crystal. Angew. Chem. 2011, 123,
442–446. [CrossRef]
31. Jahan, M.; Liu, Z.; Loh, K.P. A Graphene oxide and copper-centered metal organic framework composite as a
tri-functional catalyst for HER, OER, and ORR. Adv. Funct. Mater. 2013, 23, 5363–5372. [CrossRef]
32. Cheng, F.; Zhang, T.; Zhang, Y.; Du, J.; Han, X.; Chen, J. Enhancing electrocatalytic oxygen reduction on
MnO2 with vacancies. Angew. Chem. Int. Ed. 2013, 52, 2474–2477. [CrossRef] [PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

View publication stats

You might also like